You are on page 1of 9

Applied Catalysis A: General 301 (2006) 16–24

www.elsevier.com/locate/apcata

Comparison of reducibility and stability of alumina-supported


Ni catalysts prepared by impregnation and co-precipitation
Guohui Li 1, Linjie Hu, Josephine M. Hill *
Department of Chemical and Petroleum Engineering, University of Calgary, Calgary, Alta., Canada T2N 1N4
Received 28 July 2005; received in revised form 8 November 2005; accepted 11 November 2005

Abstract
Three Ni–Al2O3 catalysts, with nickel loadings of 10–13 wt.%, were prepared by co-precipitation (Ni–Al co-precip), impregnation on an in-
house sol–gel derived alumina (Ni/sol–gel Al2O3), and impregnation on a commercial g-Al2O3 (Ni/g-Al2O3). The catalysts were characterized by
N2 physisorption, H2 chemisorption, TPR, XRD, SEM and TEM. The Ni species, Ni particle size, and Ni reducibility depended on the preparation
method. The Ni–Al co-precip and Ni/sol–gel Al2O3 catalysts contained NiAl2O4 species after calcination, while the Ni/g-Al2O3 catalyst contained
NiO and NiAl2O4 species after calcination. Although the Ni/g-Al2O3 catalyst was the easiest to reduce, according to TPR, this catalyst had the
lowest hydrogen uptake over a 100 h temperature-staged reduction experiment. The Ni–Al co-precip and Ni/sol–gel Al2O3 catalysts had Ni
dispersions of over 7% with reduction at 550 8C for 31 h, and maximum dispersions of 10%, after reduction at 650 8C for 7 h. After reduction at
550 8C, the Ni particles were not evident by TEM examination. The results suggest that the formation of a surface NiAl2O4 spinel phase during
preparation is beneficial for a high Ni dispersion in the reduced catalyst.
# 2005 Elsevier B.V. All rights reserved.

Keywords: Nickel catalysts; Alumina; Reducibility; Co-precipitation; Impregnation; Catalyst preparation; Stability; Sol–gel synthesis

1. Introduction temperature reforming processes, such as the reforming of


methanol to hydrogen for fuel cell applications [21]. At low
Reforming is a very important industrial process. The temperatures, sintering is less likely. The nickel, however, may
catalysts for the reforming of methane have been extensively be difficult to reduce. Thus, the focus of this work is to produce
studied because of their applications in the production of a well-dispersed alumina-supported nickel catalyst that can be
syngas and hydrogen (for example, see [1–6]). Most reduced at temperatures below 600 8C.
commercial nickel-based catalysts, however, suffer deactiva- In order to achieve high nickel dispersion, many researchers
tion due to sintering and carbon deposition during operation [7– have focused on the development of high surface area nickel–
11]. The nickel particle size, in particular particles larger than alumina (Ni–Al2O3) catalysts [22–25]. A variety of preparation
5–12 nm, has been considered as one of the main factors methods, such as impregnation of nickel on high surface area
influencing carbon formation [7,12–15]. The sintering of nickel alumina supports [20,26–34], nickel–alumina co-precipitation
particles not only reduces nickel utilization, but also [33,35–43], and sol–gel methods [15,17,25,33,44–47] have also
deteriorates the reaction performance of the catalysts by been studied. The interaction between nickel and the alumina
promoting carbon formation. Therefore, the objective of novel support is complex. During preparation Ni2+ could disperse on
catalyst development is to achieve both high nickel dispersion the surface of Al2O3, or diffuse into the bulk structure of Al2O3 to
through small nickel particles and long-term thermal stability different extents, depending on the intrinsic properties of the
[16–20]. Recently there has been increased interest in lower support, nickel loading, and preparation parameters. Nickel can
be highly dispersed on the surface if nickel is incorporated into
the spinel structure of alumina to form nickel aluminate
(NiAl2O4). The dispersion is maintained because of the strong
* Corresponding author. Tel.: +1 403 210 9488; fax: +1 403 284 4852.
E-mail address: jhill@ucalgary.ca (J.M. Hill).
interaction between Ni2+ and the support. The Ni2+ in NiAl2O4,
1
Present address: SNC Lavalin Inc., 909 5th Avenue SW, Calgary, Alta., however, is difficult to reduce at temperatures below approxi-
Canada T2P 3H5. mately 700 8C [32,48] and, thus, may not be active for all
0926-860X/$ – see front matter # 2005 Elsevier B.V. All rights reserved.
doi:10.1016/j.apcata.2005.11.013
G. Li et al. / Applied Catalysis A: General 301 (2006) 16–24 17

reactions, depending on the reduction temperature. High aluminium nitrate (Al(NO3)39H2O, Aldrich) was stirred at
temperature reductions tend to result in sintering. More weakly 60 8C while a 5% NH3H2O solution was added drop wise to
bound nickel species, such as nickel oxide (NiO), are easier to increase the pH to a value of 9. The temperature was then
reduce but also more easily sintered, resulting in larger particles increased to 90 8C and the solution was stirred for 1 h before
that promote carbon formation. Rather than being completely being filtered and washed five times with deionized water. This
incorporated into the alumina structure, a surface nickel spinel catalyst will be referred to as Ni–Al co-precip.
phase may also form [49]. This phase is between bulk nickel Nickel catalysts were also prepared by impregnation on two
aluminate and nickel oxide in terms of interaction with the different supports. The first support was prepared in-house
support, and may be a suitable compromise with respect to through a sol–gel method. A 1.0 M aqueous solution of
reducibility and stability, but may also be more difficult to Al(NO3)39H2O (Aldrich) and a 10% NH3 solution were added
produce. drop wise simultaneously into a beaker of deionized water heated
The preparation method strongly influences the character- to 50 8C. A precipitate formed after the pH reached a value of 9.
istics of the final catalyst. Sol–gel derived alumina materials The precipitate was aged for 10 min at 60 8C and at pH 9, then
have high surface areas, controllable nano-particle size, and filtered and washed five times with deionized water. After
versatile structures [50–52]. Nickel supported on sol–gel washing, the filter cake was placed into a beaker of deionized
derived Al2O3 catalysts have a more stable support structure water, to which a 1.0% HNO3 solution was added drop wise at
and a higher nickel dispersion than catalysts made by ambient temperature until the pH was equal to 5. The mixture was
impregnation on g-Al2O3, resulting in higher activity and stirred for 12 h. Then 10% NH3 solution was added into the sol–
resistance to carbon deposition [12,15,25]. The role of the Al2O3 solution with stirring at 50 8C until the pH increased to 9,
different phases of Ni (Ni spinel, NiO, etc.), however, is still not in order to form a gel. The filter cake was washed once with
well understood. deionized water, and then dried at 120 8C overnight, and calcined
In the present work, we examine the reducibility and stability at 600 8C for 3 h in oxygen flowing at 20 ml/min. Impregnation
of nickel catalysts supported on g-Al2O3. Three g-Al2O3- was performed by mixing this calcined sol–gel Al2O3 with an
supported nickel catalysts were prepared by co-precipitation, aqueous nickel nitrate solution (Ni(NO3)26H2O, Aldrich) at
impregnation on an in-house prepared alumina, and impregna- room temperature, followed by heating to evaporate excess
tion on a commercial g-Al2O3. The in-house prepared alumina water. This catalyst will be referred to as Ni/sol–gel Al2O3.
was made using sol–gel synthesis. The catalysts were The second support was g-Al2O3 (Alfa, activated, neutral,
characterized by N2 physisorption, H2 chemisorption, TPR, 60 mesh (0.2 mm)). The support was mixed with deionized
XRD, SEM and TEM. The effects of preparation methods on the water to wet the support and then mixed with an aqueous nickel
structure of the catalysts, Ni2+ reducibility, and Ni0 dispersion nitrate solution (Ni(NO3)26H2O, Aldrich). This slurry was
were examined. In addition, to investigate the stability of the heated to 80 8C with stirring until most of the water had
produced nickel structures, the catalysts were reduced in stages evaporated to obtain a solid. This catalyst will be referred to as
between 550 and 850 8C for a total duration of 100 h. Ni/g-Al2O3.
All Ni samples were dried at 120 8C overnight, and calcined
2. Experimental at 600 8C for 3 h in flowing oxygen (20 ml/min). The Ni
loadings of the calcined samples were obtained by ICP-AES
2.1. Catalyst preparation measurement (Galbraith Laboratories Inc.) and are reported in
Table 1.
Nickel–alumina catalysts were prepared using co-precipita-
tion or impregnation. The co-precipitation method involved 2.2. Catalyst characterization
mixing nickel and aluminium precursor solutions, and then
adjusting the pH to precipitate a solid. Specifically, an aqueous The N2 adsorption–desorption isotherms for the catalysts
solution of nickel nitrate (Ni(NO3)26H2O, Aldrich) and were measured on a Quantachrome instrument AUTOSORB-1-

Table 1
Properties of Ni–Al2O3 catalysts before and after reduction at various temperatures
Catalyst Ni loading (%) Surface areaa (m2/g) Pore volumea (ml/g)
Calcined Reduced at Reduced at 850 8C c Calcined Reduced at
550 8C b 850 8C c
Ni–Al co-precip 12 373 350 160 0.96 0.87
Sol–gel Al2O3 – 303 1.3
Ni/sol–gel Al2O3 10 232 208 199 1.0 1.0
g-Al2O3 – 180 0.30
Ni/g-Al2O3 13 141 118 98 0.22 0.24
a
Error is 5% in the surface area and the pore volume measurements.
b
Reduction time was 1 h at 550 8C.
c
Reduction time was 31 h at 850 8C as part of the temperature-staged reduction.
18 G. Li et al. / Applied Catalysis A: General 301 (2006) 16–24

C. All samples were evacuated at 400 8C until the outgas rate reduced completely, and the Ni particles were spherical, the Ni
was below 15 mm Hg/min (or 2 Pa/min) prior to the measure- particle diameter, dNi = 6V/S, where V is the volume of total
ment. The specific surface area was calculated using the BET metallic Ni, and S is the active Ni surface area, was calculated.
method. The total pore volume was determined at a relative An adsorption stoichiometry of one hydrogen atom adsorbed
pressure P/P0 = 0.99. Pore size distributions were calculated per surface nickel atom (H/Nis = 1) was assumed. The percent
from the desorption isotherms using the BJH method. Ni dispersion is calculated by dividing the number of exposed
Powder X-ray diffraction (XRD) spectra were recorded on a surface Ni atoms (as determined by H2 chemisorption) by the
Rigaku Multiflex X-ray diffractometer using Cu Ka1 radiation total amount of Ni in the catalyst.
(l = 1.54056 Å) at 40 kV tube voltage and 40 mA tube current
with a scanning speed of 28 min 1 between 158 and 808 2u. The 3. Results and discussion
XRD patterns were referenced to the powder diffraction files
(ICDD-FDP database) for identification. The average crystal- 3.1. Characterization of the calcined catalysts
lite diameter of metallic Ni was calculated using the Scherrer
method, DNi = Kl/b cos u, where the constant K was taken as The surface areas and pore volumes determined by N2
0.9, b was the full width at half maximum (FWHM) of the Ni physisorption for the Ni–Al2O3 catalysts are given in Table 1.
(2 0 0) peak at 2u = 51.88. The isotherms and pore size distributions are shown in Fig. 1. A
Scanning electron microscopy (SEM) images were recorded value of 180 m2/g for the commercial g-Al2O3 support
on a Philips/FEI ESEM. The sample was coated with gold before compares well with the nominal value of 150 m2/g given by
the measurement. Transmission electron microscopy (TEM) the manufacturer. The in-house prepared sol–gel alumina
images were recorded on an H-7000 transmission electron support had a significantly higher surface area of 303 m2/g but
microscope (Hitachi) at 75 kV. The samples were ground to a fine also had a significantly wider pore size distribution than the
powder, and put into acetone to make a suspension. A drop of the commercial g-Al2O3 (Fig. 1).
suspension was placed on a lacey carbon grid. Addition of Ni onto these supports reduced the surface areas
Temperature-programmed reduction (TPR) of the samples and pore volumes, and shifted the pore size distributions to
was performed on a CHEMBET 3000 apparatus (Quanta- smaller pore diameters, suggesting that the impregnated Ni
chrome) with a thermal conductivity detector (TCD). For this blocks some pores in the support. The surface areas for the Ni
measurement, approximately 100 mg of sample was placed into catalysts varied between 141 and 373 m2/g, while the pore
a U-shaped quartz tube (ID = 4 mm), and reduced in a flow of volumes varied between 0.22 and 1.0 ml/g. The surface area
reducing gas (5% H2/N2) at 15 ml/min. The sample was heated and pore volume were the smallest for the Ni/g-Al2O3. The Ni–
from room temperature to 1000 8C with a heating rate of 10 8C/ Al co-precip catalyst had the largest surface area, while the Ni/
min. The reduction extent of the catalysts was estimated by sol–gel Al2O3 catalyst had the largest pore volume. The Ni–Al
comparing the theoretical hydrogen consumption to the actual co-precip catalyst had a bimodal pore size distribution with one
hydrogen consumption. The TCD signal was calibrated by peak around 3 nm and another larger broader peak around
injecting a known volume of H2 and measuring the peak area. 9 nm. In comparison, the peak maximums occurred at

2.3. Temperature-staged reduction

The temperature-staged reduction experiment involved


measuring the hydrogen chemisorption of the three catalysts
after sequential reductions at temperatures between 550 and
850 8C. The total reduction time at each temperature was 31 h,
with the exception of 650 8C, at which temperature the total
reduction time was 7 h. The total reduction time for the entire
experiment was 100 h, and the samples were not exposed to air
during this time but were kept under hydrogen or in vacuum on
the chemisorption instrument.
H2 chemisorption measurements were carried out on the
same apparatus used for the N2 adsorption (AUTOSORB-1-C,
Quantachrome). For these measurements, approximately 1.0 g
of catalyst was placed in a U-shaped quartz tube (ID = 10 mm),
and reduced in a H2 flow of 15 ml/min at temperatures of 550,
650, 750 or 850 8C for times between 1 and 12 h. After
reduction, the sample was evacuated, at the same temperature
as reduction, for 2 h, then cooled down to 40 8C for the H2
chemisorption measurement. The H2 monolayer uptake of the Fig. 1. Adsorption/desorption isotherms (a) and pore size distributions (b) of
catalysts was calculated by extrapolating the H2 adsorption supports and Ni catalysts (open symbols – support, closed symbols – Ni/
isotherm to zero pressure. Assuming that Ni2+ ions were support).
G. Li et al. / Applied Catalysis A: General 301 (2006) 16–24 19

Fig. 2. SEM images of Ni–Al2O3 catalysts after calcination: (a) Ni–Al co-precip, (b) Ni/sol–gel Al2O3, and (c) Ni/g-Al2O3.

approximately 8 and 4 nm for the Ni/sol–gel Al2O3 and Ni/g- The other peaks for the defect spinel phase occur at values of
Al2O3 catalysts, respectively. SEM images representative of the approximately 2u = 378, 458, 608 and 668 [17]. The diffraction
morphology of the catalysts in their as produced state are shown angles for defect spinel shift according to the nickel content and
in Fig. 2. The particle sizes are significantly smaller for the Ni– preparation variables [17,54]. The first two peaks overlap with
Al co-precip and Ni/sol–gel Al2O3 catalysts than the Ni/g- the peaks for the alumina support, although the relative intensity
Al2O3 catalyst. of these peaks has increased. Kim et al. [17] obtained the same
The XRD spectra for the calcined Ni–Al2O3 catalysts are XRD result for defect Ni–Al spinel structure. The authors
shown in Fig. 3. The spectrum for g-Al2O3 has a spinel structure suggested that the lattice parameters of the impregnated Ni–
[17] and is shown for comparison. The spectrum for Ni/g-Al2O3 Al2O3 catalyst increased due to the Ni2+ incorporated into the g-
is similar to that for the support except for additional peaks at 2u Al2O3 lattice. Since the ionic radius of nickel is larger than that of
values of 438 and 638 that are associated with NiO [32,48]. In aluminium, the incorporation of nickel expands the lattice and
contrast, NiO is not observed on the other two catalysts. XRD the diffraction peak shifts. That is, the NiAl2O4 spinel has a larger
can typically only detect metal crystallites that are larger than lattice parameter than that of g-Al2O3, and hence, the diffraction
2–5 nm [53] and, thus, it is possible that NiO exists on these peak for the (4 4 0) plane of NiAl2O4 spinel has a smaller 2u value
catalysts if the crystallite sizes are smaller than the detection at 2u = 668, compared with 2u = 67.38 for g-Al2O3. There may be
limit. a slight shift in this peak for the Ni/g-Al2O3 catalyst. The relative
The peak at 2u = 668 for the Ni–Al co-precip and Ni/sol–gel intensity of the peak around 378 is consistent with NiAl2O4 being
Al2O3 catalysts is between the values of 2u = 67.38 for g-Al2O3 present on this catalyst.
and 2u = 65.58 for stoichiometric NiAl2O4 spinel, suggesting
that a defect NiAl2O4 phase has been formed on these catalysts. 3.2. Nickel reducibility – TPR and TSR

In order to investigate the reducibility of Ni2+ ions in the Ni–


Al2O3 catalysts and their stability, temperature-programmed
reduction (TPR) and temperature-staged reduction (TSR), were
used to study the samples. The catalysts were characterized by
XRD, TEM, and N2 physisorption after both reduction
procedures.
The TPR profiles for the three catalysts are shown in Fig. 4.
The results for g-Al2O3 and NiO are included for comparison.
The extent of reduction of the Ni in the catalysts after TPR to
1000 8C was determined to be greater than 92% for all of the
catalysts, indicating that essentially all of the Ni in the catalysts
is reduced in the TPR experiments.
No reduction peaks were observed on the g-Al2O3 sample,
as expected. The NiO sample has a reduction peak around
385 8C, which agrees well with the literature [55]. The Ni/g-
Al2O3 catalyst has a peak at 410 8C, which indicates the
presence of NiO on the Ni/g-Al2O3 catalyst. All other reduction
peaks were above 600 8C, as is typical in Ni/alumina catalysts
containing Ni spinel [17]. Thus, the TPR results are consistent
with the XRD analysis (Fig. 3) in that only Ni/g-Al2O3 had an
XRD peak corresponding to NiO, while all catalysts had XRD
peaks corresponding to NiAl2O4.
The Ni–Al co-precip catalyst was reduced between 650 and
1000 8C. The maximum of the reduction peak occurs around
Fig. 3. X-ray diffraction spectra of Ni–Al2O3 catalysts after calcination. 870 8C, which is consistent with the literature [28,32]. The
20 G. Li et al. / Applied Catalysis A: General 301 (2006) 16–24

The behaviour of both Ni–Al co-precipitated and Ni/sol–gel


Al2O3 catalysts during the temperature-staged reduction
procedure was similar. The H2 uptake increased with
temperature and time to a maximum of approximately
49 mmol/mol Ni after reduction for 38 h. The H2 uptake then
decreased over the next 62 h of reduction to approximately
31 mmol/mol Ni after 100 h of reduction. The maximum H2
uptake of these catalysts is three times higher than that of the
Ni/g-Al2O3 catalyst with dispersions slightly less than 10%.
The decrease in H2 uptake for both Ni–Al co-precip and Ni/sol–
gel Al2O3 catalysts compared, to that of the Ni/g-Al2O3
catalyst, is consistent with the former catalysts having smaller
metallic nickel particles that are more easily sintered than the
larger particles on the latter catalyst [9].

3.3. Characterization after reduction – Ni stability

Representative TEM images of the catalysts after reduction


at 550 8C for 1 h, and after reduction at 850 8C for 31 h at the
end of the temperature-staged reduction experiment are shown
Fig. 4. Temperature-programmed reduction profiles of Ni–Al2O3 catalysts; in Fig. 6. The TEM images of the samples after TPR are very
heating rate of 10 8C/min in 5% H2 in N2.
similar to those after temperature-staged reduction. After 1 h of
reduction, it is difficult to unambiguously discern Ni particles
Ni/sol–gel Al2O3 catalyst was reduced between 500 and on any of the catalysts (Fig. 6a–c). Note, however, that the
950 8C, with a peak maximum around 800 8C, while the Ni/g- contrast between surface Ni particles and the support is reduced
Al2O3 catalyst was reduced between 500 and 920 8C, with a because Ni has been incorporated into the alumina support
peak maximum around 760 8C. Therefore, Ni2+ is easiest to structure as spinel. After 100 h of reduction (Fig. 6d–f), Ni
reduce in Ni/g-Al2O3, more difficult to reduce in Ni/sol–gel particles are evident on all the catalysts. The particles are less
Al2O3, and most difficult to reduce in Ni–Al co-precip catalyst. than 20 nm on the Ni/sol–gel Al2O3 and Ni–Al co-precip
The change in hydrogen uptake with reduction time and catalysts, while the Ni/g-Al2O3 catalyst has larger particles
temperature was studied with temperature-staged reduction. ranging up to 40 nm in size.
The results are shown in Fig. 5. The Ni/g-Al2O3 catalyst had the The growth of the Ni particles is also evident in the XRD
lowest H2 uptake with a maximum uptake of 15 mmol/mol Ni spectra of the reduced catalysts (Figs. 7–9). Fig. 7 shows the
after reduction at 650 8C for 1 h. This uptake corresponds to a XRD spectra of the Ni–Al co-precip catalyst from calcination
Ni dispersion of less than 3%. The H2 uptake does not change through to after TPR. After reduction at 550 8C for 1 h the XRD
substantially with time at 750 and 850 8C, indicating that the spectra is essentially the same as that of the calcined catalyst.
reduced Ni0 phase is stable, or the rate of Ni2+ reduction and Ni0 After TSR and TPR, however, peaks are present at 2u = 44.58,
sintering are comparable. 528, and 768 indicating that metallic Ni has been formed. At the
same time, a peak shift from 2u = 66.08 (assigned to defect
NiAl2O4 spinel) to 2u = 67.38 (assigned to Al2O3) has occurred,
suggesting that the Ni2+ in the spinel phase has been reduced to
Ni0. The decrease in the relative intensity of the Al2O3 peaks
compared to the Ni peaks indicates that the surface of the
catalyst has become enriched in Ni. The XRD spectra of the Ni/
sol–gel Al2O3 catalyst (Fig. 8) are very similar to those of the
Ni–Al co-precipitated catalyst in terms of the change in the
oxidation state of the Ni between calcination and reduction.
The relative intensity of the Ni peaks is higher after TPR to
1000 8C than after temperature-staged reduction to 850 8C, for
both catalysts, indicating that temperature has a greater
influence on reduction than time.
The Ni/g-Al2O3 catalyst (Fig. 9) is different than the other
two catalysts. After calcination, NiO was present on the catalyst
in addition to a NiAl2O4 phase. After reduction, the NiO peaks
at 2u = 438 and 638 disappeared, and the Ni0 peaks at
2u = 44.58, 528, and 768 appeared. Thus, as expected, the
Fig. 5. H2 uptake of Ni/alumina catalysts during temperature-staged reduction. NiO phase is reduced easily to produce metallic Ni.
G. Li et al. / Applied Catalysis A: General 301 (2006) 16–24 21

Fig. 6. TEM images of Ni/alumina catalysts after reduction: (a) Ni–Al co-precip reduced at 550 8C; (b) Ni/sol–gel Al2O3 reduced at 550 8C; (c) Ni/g-Al2O3 reduced
at 550 8C; (d) Ni–Al co-precip after TSR; (e) Ni/sol–gel Al2O3 after TSR; (f) Ni/g-Al2O3 after TSR.

3.4. Stability of catalyst structure decreased less than 10% for the Ni–Al co-precip catalyst, while
the pore volumes of the other catalysts remained stable.
Table 1 shows the change of surface area for all catalysts Fig. 10 shows the change in pore size distribution of the
after reduction. After reduction at 550 8C for 1 h, the surface catalysts after calcination and after the temperature-staged
area decreased approximately 10% for all the catalysts. After reduction experiments. In all cases, the mean pore sizes
temperature-staged reduction, the surface areas decreased by increased after 100 h of reduction. The Ni–Al co-precip catalyst
approximately 60%, 30% and 15% for the Ni–Al co-precip, Ni/ changed the most significantly with the maximum pore size
g-Al2O3, and Ni/sol–gel Al2O3 catalysts, respectively. The approximately doubling. These results are in agreement with
higher surface area of the Ni/sol–gel Al2O3 catalyst after the the changes in the surface areas and pore volumes (Table 1).
temperature-staged reduction indicates that the structural Significant changes in the structure of the sol–gel Al2O3 and
stability of this catalyst is good at 850 8C. The pore volume g-Al2O3 supports are not evident in the TEM images (Fig. 6).
22 G. Li et al. / Applied Catalysis A: General 301 (2006) 16–24

Fig. 9. X-ray diffraction spectra of the Ni/g-Al2O3 catalyst after various


Fig. 7. X-ray diffraction spectra of Ni–Al co-precip catalyst after various treatments.
treatments.

Some needle-like structures were observed on the Ni/sol–gel


Al2O3 catalyst after temperature-staged reduction. Experiments
were done with the sol–gel support not impregnated with Ni.
The needle-like structures appeared after a similar heat
treatment to that experienced in the reduction experiment.

Fig. 8. X-ray diffraction spectra of the Ni/sol–gel Al2O3 catalyst after various Fig. 10. Pore size distribution of Ni–Al2O3 catalysts before and after tempera-
treatments. ture-staged reduction.
G. Li et al. / Applied Catalysis A: General 301 (2006) 16–24 23

Thus, the structures are related to the alumina and not the Ni. Table 2
Ni particle diameters after reduction by temperature-staged reduction (TSR) or
The pore structure of the Ni–Al co-precip catalyst is less sharp
temperature-programmed reduction (TPR)
in the TEM images after reduction, which is consistent with a
significant change in the pore size distribution. Additional Catalyst Ni particle diameter (nm)
peaks, around 2u = 32–338 are evident in the spectra of the Ni– After TSR After TPR
Al co-precip catalyst after TSR and TPR (Fig. 7). This change XRD TEM H2 uptake XRD
may be due to the conversion of the support to u-Al2O3 [56].
Additional peaks were also evident in the XRD spectra of the Ni–Al co-precip 11 10 16 15
Ni/sol–gel Al2O3 12 10–20 17 15
Ni/g-Al2O3 after TPR at 2u = 388 and 788, which may also be Ni/g-Al2O3 15 30 40 15
related to a phase change.

3.5. Comparison of preparation methods sites because of a higher binding energy [61]. We can speculate
that the structure of the sol–gel support is more conducive to the
In this work, preparation using either co-precipitation or formation of Ni in octahedral sites on the alumina and for this
impregnation on an in-house prepared sol–gel alumina resulted reason, the TPR spectra of the Ni/sol–gel Al2O3 catalyst is
in Ni particles that were less than 15 nm in size. Our results shifted to lower temperatures relative to the Ni–Al co-precip
indicate that a high surface area support is not sufficient for high catalyst (Fig. 4). In addition, the sol–gel support was more
Ni dispersion. That is, the Ni–Al co-precip catalyst had a stable than the co-precipitated support. After reduction at
significantly higher surface area than the Ni/sol–gel Al2O3 850 8C, the co-precipitated support had significantly larger
catalyst, and yet both catalysts had similar H2 uptakes and Ni pores (Fig. 10) and a significantly decreased surface area
dispersions. The impregnation of Ni onto the g-Al2O3 support (Table 1).
resulted in a NiO phase and a NiAl2O4 phase (Figs. 3 and 4). In The stability of the Ni particles is similar on the Ni–Al co-
contrast, on the other catalysts, only a NiAl2O4 phase was precip and Ni/sol–gel Al2O3 catalysts (Fig. 7). Table 2
formed, and it was this defect spinel phase that was reduced to summarizes the Ni particle size measured by different
form metallic Ni. Although the spinel phase is more difficult to techniques after TPR or temperature-staged reduction. The
reduce than NiO, the Ni–Al co-precip and Ni/sol–gel Al2O3 results from TEM and H2 uptake are similar within the errors of
catalysts had three times greater H2 uptakes (Fig. 5) than the Ni/ each technique. The limitation of TEM is the difficulty
g-Al2O3 catalyst. These results suggest that a surface, rather distinguishing the Ni particles from the support, which contains
than bulk, spinel phase was formed. Ni atoms in the bulk or surface spinel structure; while, the
The Ni/sol–gel Al2O3 catalyst was easier to reduce at 550 8C limitation with H2 uptake is the assumption that all of the Ni is
than the Ni–Al co-precip catalyst (Fig. 4), which may be reduced. It is expected that XRD will give the smallest sizes
important for preventing sintering of the Ni particles. The Ni on because XRD detects the crystal size, which could be smaller
the co-precipitated catalyst may have been more difficult to than the particle size (i.e., the particles could be polycrystalline
reduce because of more intimate mixing between the Ni and the or be amorphous). In fact, after TPR, the Ni crystallite size was
alumina resulting from the preparation method. In contrast, the identical on all of the catalysts. The Ni–Al co-precip and Ni/
Ni on the sol–gel support is not as incorporated into the sol–gel Al2O3 catalysts have similar Ni particle sizes up to
structure. Studies on alumina-supported zinc catalysts [57], 20 nm after reduction, while the Ni/g-Al2O3 had the largest Ni
silica-supported titania [58] both have shown that sol–gel particle sizes up to 40 nm. The Ni particle sizes are comparable
preparation methods produce more dispersed metal catalysts to those for Ni/a-alumina of 0.4–2.6 wt.% Ni [62] and Ni/g-
than other traditional preparation methods (precipitation or Al2O3 of 7–20 wt.% [15] produced by sol–gel synthesis. The
wet-mixing). In addition, the catalyst supports have more next step in our studies is to test and compare these catalysts for
favourable properties, including mono-modal pore size reforming reactions.
distributions, and high pore volumes, when the catalysts are
prepared by sol–gel methods [57,58], which is consistent with 4. Conclusions
the results in this study.
It has been shown that the Ni precursor, Ni loading and In this work, we have prepared well-dispersed Ni/alumina
calcination temperature can influence the characteristics of catalysts that were reducible at 550 8C. Three preparation
alumina-supported Ni catalysts [17,53]. In this study, the same methods – co-precipitation, impregnation on an in-house
Ni precursor (Ni nitrate), Ni loading (11 wt.%) and the same prepared sol–gel alumina, and impregnation on a commercial
calcination conditions (600 8C for 3 h in flowing oxygen) have g-Al2O3 – were compared. The first two preparations resulted
been used for all of the catalysts. The localized heating during in catalysts with Ni particles less than 5 nm in size after
calcination may be different for the three catalysts because of reduction at 550 8C. These particles sintered after reduction at
the differences in the support structures, possibly resulting in 850 8C to particles 15 nm in size. Based on XRD analysis,
different Ni species and particle sizes. X-ray photoelectron surface NiAl2O4 phases were produced on the Ni–Al co-precip
spectroscopy (XPS) has been used to examine the location of Ni and Ni/sol–gel Al2O3 catalysts, while NiO and bulk NiAl2O4
species in the alumina matrix [59–61]. Ni located in tetrahedral phases were produced on the Ni/g-Al2O3 catalyst. Although,
sites is more difficult to reduce than Ni located in octahedral NiO is easier to reduce than NiAl2O4, the Ni–Al co-precip and
24 G. Li et al. / Applied Catalysis A: General 301 (2006) 16–24

Ni/sol–gel Al2O3 catalysts had significantly higher H2 uptakes [25] S. Tang, L. Ji, J. Lin, H.C. Zeng, K.L. Tan, K. Li, J. Catal. 194 (2000) 424.
[26] Y.H. Zhang, G.X. Xiong, S.S. Sheng, S.L. Liu, W.S. Yang, Acta Phys.-
than the Ni/g-Al2O3 catalyst. The maximum dispersions
Chim. Sin. 15 (1999) 735.
achieved on the former two catalysts were 9.5% after reduction [27] S. Wang, G.Q.M. Lu, Appl. Catal. B 16 (1998) 269.
at 650 8C for 7 h. The dispersions decreased to 6% after [28] Z.X. Cheng, X.G. Zhao, J.L. Li, Q.M. Zhu, Appl. Catal. A 205 (2001) 31.
reduction at 850 8C for 31 h (a total reduction time of 100 h). [29] Z. Xu, Y. Li, J. Zhang, L. Chang, R. Zhou, Z. Duan, Appl. Catal. A 210
The Ni/g-Al2O3 catalyst maintained a dispersion of 3% during (2001) 45.
the 100 h temperature-staged reduction, and had Ni particles [30] C. Courson, L. Udron, C. Petit, A. Kiennemann, Sci. Tech. Adv. Mater. 3
(2002) 271.
of 20–40 nm in size. The sol–gel derived Al2O3 structure was [31] S. Takenaka, H. Ogihara, I. Yamanaka, K. Otsuka, Appl. Catal. A 217
the most stable support up to temperatures of 1000 8C. (2001) 101.
[32] H.-S. Roh, K.-W. Jun, S.-E. Park, Appl. Catal. A 251 (2003) 275.
Acknowledgements [33] A.I. Tsyganok, T. Tsunoda, S. Hamakawa, K. Suzuki, K. Takehira, T.
Hayakawa, J. Catal. 213 (2003) 191.
[34] J. Guo, H. Lou, H. Zhao, D. Chai, X. Zheng, Appl. Catal. A 273 (2004) 75.
We would like to acknowledge the Natural Sciences and [35] J.-M. Wei, B.-Q. Xu, J.-L. Li, Z.-X. Cheng, Q.-M. Zhu, Appl. Catal. A 196
Engineering Research Council (NSERC) for funding this (2000) L167.
project. We thank Dr. P.R. Pereira for allowing us to use his [36] Y. Cesteros, P. Salagre, F. Medina, J.E. Sueiras, Chem. Mater. 12 (2000)
ChemBET instrument, Mr. R. Humphrey at the Microscopy and 331.
Imaging Facility, University of Calgary, for assistance with [37] Z. Hou, T. Yashima, Appl. Catal. A 261 (2004) 205.
[38] G.-J. Li, X.-X. Huang, M. Ruan, J.-K. Guo, Ceram. Int. 28 (2002) 165.
both SEM and TEM analysis, and Ms. L. Klatzel-Mudry for [39] M.A. Marturano, E.F. Aglietti, O.A. Ferretti, Thermochim. Acta 336
assistance with the XRD analysis. (1999) 47.
[40] K. Takehira, T. Shishido, P. Wang, T. Kosaka, K. Takaki, J. Catal. 221
References (2004) 43.
[41] R. Villa, C. Cristiani, G. Groppi, L. Lietti, P. Forzatti, U. Cornaro, S.
[1] D.J. Wilhelm, D.R. Simbeck, A.D. Karp, R.L. Dickenson, Fuel Proc. Tech. Rossini, J. Mol. Catal. A: Chem. 204/205 (2003) 637.
[42] T. Shishido, M. Sukenobu, H. Morioka, R. Furukawa, H. Shirahase, K.
71 (2001) 139.
[2] J.R. Rostrup-Nielsen, Catal. Today 18 (1993) 305. Takehira, Catal. Lett. 73 (2001) 21.
[3] A.K. Avci, Z. Ilsen Onsan, D.L. Trimm, Appl. Catal. A 216 (2001) 243. [43] E. Kis, R. Marinkovic-Neducin, G. Lomic, G. Boskovic, D.Z. Obadovic, J.
[4] Z.-W. Liu, K.-W. Jun, H.-S. Roh, S.-E. Park, J. Power Sources 111 (2002) Kiurski, P. Putanov, Polyhedron 17 (1998) 27.
[44] C. Otero Arean, M. Penarroya Mentruit, A.J. Lopez Lopez, J.B. Parra,
283.
[5] C. Song, Catal. Today 77 (2002) 17. Coll. Surf. A: Phys.-Chem. Eng. Aspects 180 (2001) 253.
[6] J. Sun, X. Qiu, F. Wu, W. Zhu, W. Wang, S. Hao, Int. J. Hydrogen Energy [45] D.J. Suh, T.-J. Park, J.-H. Kim, K.-L. Kim, J. Non-Cryst. Solids 225 (1998)
29 (2004) 1075. 168.
[46] Z. Xu, Y. Li, J. Zhang, L. Chang, R. Zhou, Z. Duan, Appl. Catal. A 213
[7] J.R. Rostrup-Nielsen, J. Sehested, J.K. Norskov, Adv. Catal. 47 (2002) 65.
[8] J.R. Rostrup-Nielsen, Catal. Today 63 (2000) 159. (2001) 65.
[9] J. Sehested, J.A.P. Gelten, I.N. Remediakis, H. Bengaard, J.K. Norskov, J. [47] S. Krompiec, J. Mrowiec-Bialon, K. Skutil, A. Dukowicz, L. Pajak, A.B.
Jarzebski, J. Non-Cryst. Solids 315 (2003) 297.
Catal. 223 (2004) 432.
[10] A.C.S.C. Teixeira, R. Giudici, Chem. Eng. Sci. 56 (2001) 789. [48] C. Li, Y.-W. Chen, Thermochim. Acta 256 (1995) 457.
[11] S.C. Tsang, J.B. Claridge, M.L.H. Green, Catal. Today 23 (1995) 3. [49] M. Lo Jacono, M. Schiavello, A. Cimino, J. Phys. Chem. 75 (1971) 1044.
[12] Y. Zhang, G. Xiong, S. Sheng, W. Yang, Catal. Today 63 (2000) 517. [50] X. Bokhimi, J. Sanchez-Valente, F. Pedraza, J. Solid State Chem. 166
(2002) 182.
[13] J. Sehested, A. Carlsson, T.V.W. Janssens, P.L. Hansen, A.K. Datye, J.
Catal. 197 (2001) 200. [51] U. Janosovits, G. Ziegler, U. Scharf, A. Wokaun, J. Non-Cryst. Solids 210
[14] F.B. Rasmussen, J. Sehested, H.T. Teunissen, A.M. Molenbroek, B.S. (1997) 1.
Clausen, Appl. Catal. A 267 (2004) 165. [52] F. Mange, D. Fauchadour, L. Barre, L. Normand, L. Rouleau, Coll. Surf.
A: Phys.-Chem. Eng. Aspects 155 (1999) 199.
[15] J.-H. Kim, D.J. Suh, T.-J. Park, K.-L. Kim, Appl. Catal. A 197 (2000) 191.
[16] K. Fujimoto, O. Yamazaki, K. Tomishige, Appl. Catal. A 136 (1996) 49. [53] G. Poncelet, M. Centeno, R. Molina, Appl. Catal. A 288 (2005) 232.
[17] P. Kim, Y. Kim, H. Kim, I.K. Song, J. Yi, Appl. Catal. A 272 (2004) 157. [54] X. Cai, Z. Ren, T. Hu, Y. Xie, Surf. Interf. Anal. 32 (2001) 293.
[18] J.H. Edwards, A.M. Maitra, Fuel Proc. Tech. 42 (1995) 269. [55] W. Shan, M. Luo, P. Ying, W. Shen, C. Li, Appl. Catal. A 246 (2003) 1.
[56] R.S. Zhou, R.L. Snyder, Acta Crystallogr., Sect. B: Struct. Sci. 47 (1991)
[19] M.A. Marturano, E.F. Aglietti, O.A. Ferretti, Thermochim. Acta 336
(1999) 55. 617.
[20] Y.-S. Oh, H.-S. Roh, K.-W. Jun, Y.-S. Baek, Int. J. Hydrogen Energy 28 [57] M. Valenzuela, P. Bosch, G. Aguilar-Rios, A. Montoya, I. Schifter, J. Sol–
Gel Sci. Technol. 8 (1997) 107.
(2003) 1387.
[21] R. Farrauto, S. Hwang, L. Shore, W. Reuttinger, J. Lampert, T. Giroux, Y. [58] M. Montes, F. Getton, M.S.W. Vong, P.A. Sermon, J. Sol–Gel Sci.
Liu, O. Ilinich, Annu. Rev. Mater. Res. 33 (2003) 1. Technol. 8 (1997) 131.
[22] T.V. Choudhary, C. Sivadinarayana, D.W. Goodman, Chem. Eng. J. 93 [59] P. Dufresne, E. Payen, J. Grimblot, J. Bonnelle, J. Phys. Chem. 85 (1981)
2344.
(2003) 69.
[23] C.E. Quincoces, E.I. Basaldella, S.P. De Vargas, M.G. Gonzalez, Mater. [60] M. Wu, D. Hercules, J. Phys. Chem. 83 (1979) 2003.
Lett. 58 (2004) 272. [61] C. Li, A. Proctor, D. Hercules, Appl. Spectrosc. 38 (1984) 880.
[24] A. Valentini, N.L.V. Carreno, L.F.D. Probst, E.R. Leite, E. Longo, Micro. [62] B. Li, R. Watanabe, K. Maruyama, K. Kunimori, K. Tomishige, Catal.
Today 104 (2005) 7.
Meso. Mater. 68 (2004) 151.

You might also like